Skip to main content

Full text of "Selected problems in astrophysics of compact objects"

See other formats


Selected problems in astrophysics of compact objects 



Armen Sedrakian 

Institute for Theoretical Physics, J. W. Goethe-University, 
D-60438 Frankfurt-Main, Germany 

E-mail: sedrakian@th.physik .uni-f rankf urt .de 

Abstract. I review three problems in astrophysics of compacts stars: (i) the phase diagram 
of warm pair-correlated nuclear matter at sub-saturation densities at finite isospin asymmtery; 
(ii) the Standard Model neutrino emission from superfluid phases in neutron stars within the 
Landau theory of Fermi (superfluid) liquids; (iii) the beyond Standard Model physics of axionic 
cooling of compact stars by the Cooper pair-breaking processes. 



1. Introduction 

The present lecture summarizes some of the recent work on three problems in the astrophysics 
of compact stars listed in the abstract. It attempts to provide sufficient perspective and detail 
within the limited span of pages as to be a useful introduction to these problems. The selection 
of the topics is largely motivated by the author's interests and covers only a small part of many 
facets of compact star physics. 

Neutron stars are born in spectacular explosions of type-II supernovas [Ij. The equation of 
state of matter at subnuclear densities and more generally its phase structure and composition 
are important ingredients of supernova physics, one key reason being their importance for 
the formation of neutrino signal at the neutrino-sphere and for transport of the energy that 
ultimately may generate a successful explosion. The temperature range relevant for subnuclear 
densities is several Me Vs. In Sec. [2] we will discuss some recent progress in the understanding 
of the phase diagram of nuclear matter in this regime. 

If a neutron star is formed in a supernova explosion, its further thermal evolution is 
determined by the neutrino emission from the interior of the star: neutrinos once produced 
in the star escape the stellar environment without further interactions [21 O Hj. The black- 
body radiation form the surface of the star in thermal equilibrium, while providing us with the 
information on the composition of matter and effective temperature of the emitting regions, is 
an unimportant cooling agent until the star is hundreds of thousands years old. The physics 
of neutrino emission from the interiors of neutron stars is discussed in Sec. [3] in the case of 
(non-exotic) baryonic matter. 

Non-Standard Model particles emitted during the neutrino stage of cooling of a neutron star 
can drain sufficient energy as to "spoil" the Standard Model based cooling scenario. This may 
be used to place limits on the properties of exotic particles, such as the axions. Astrophysical 
bounds on the properties of axions, i.e., the mass and the coupling to the Standard Model 
particles, have been derived from stellar (red giant, white dwarf, supernova, and neutron star) 
physics, which are complementary to those obtained from cosmology and terrestrial laboratory 
searches [5]. Section H] discusses the bounds on the axion properties derived from the cooling of 



neutron stars through inclusion of a new process of axion emission via the Cooper pair-breaking 
in superfluid baryonic matter [6]. 

2. Warm isospin asymmetrical nuclear matter 

Nuclear matter at sub-saturation densities, i.e., at p < 0.5 po, where po = 2.8 x 10 14 g cm~ 3 is 
the nuclear saturation density, is a many-body system with well defined pair interactions, whose 
dominant attractive part is responsible for the formation of nuclear clusters and the Bardeen- 
Cooper-Schrieffer (BCS) type pair-condensate(s) of nucleons. In the supernova context the 
matter is at finite, but small compared to cold neutron stars, isospin asymmetry. Furthermore, 
low densities and relatively high (again compared to neutron stars) temperatures allow for the 
existence of a substantial amount of light clusters, notably deuterons, tritons, He 3 nuclei and 
alpha particles El U QUI CED 021 03] . 

Because the unbound nucleons and deuterons are the dominant component of the matter at 
all relevant densities it is useful to consider only two-body correlations first. In the extreme low 
density limit such matter can be considered as a mixture of quasi-free nucleons and deuterons, 
where the only effect of the interaction is to renormalize the mass of the constituents. The 
deuterons being bosons may also condense and from a Bose-Einstein condensate (BEC) even 
without interactions if, of course, the temperature is sufficiently low [TH [151 US El 031 [19] - 
Imagine now increasing density while keeping the temperature of the system constant, such 
that the degeneracy of the system is effectively increased. As the density p — > po the 
abundances of deuterons are reduced because of the Pauli-blocking of the phase space available 
to nucleons [16} 120] . However, the emergence of the Fermi-surface of nucleons (at finite isospin - 
two Fermi surfaces of neutrons and of protons) leads to BCS type correlations and formation of 
macroscopic coherent state - the condensate of nucleons. The interaction channel responsible for 
this pairing phenomenon is the s Si- 3 D\ partial wave, i.e., the same interaction channel which 
binds the deuteron. Thus, it has been conjectured that the nuclear matter may undergo a BCS- 
BEC phase transition, first in the context of intermediate energy heavy-ion collisions [14^ [To] 
and more recently in the context of supernovas [SJ |2"T] . 

The theoretical framework for the description of the BCS-BEC transition was developed 
by Nozieres and Schmitt-Rink in condensed matter physics [22 1. Isospin asymmetry, induced 
by weak interactions in stellar environments and expected in exotic nuclei, disrupts isoscalar 
neutron-proton (np) pairing, since the mismatch in the Fermi surfaces of protons and neutrons 
suppresses the pairing correlations |23| . The standard Nozieres-Schmitt-Rink theory of the 
BCS-BEC crossover must also be modified, such that the low-density asymptotic state becomes 
a gaseous mixture of neutrons and deuterons [23] . 

Two relevant energy scales for the problem domain under study are provided by the shift 
5p = (p n — p p )/2 in the chemical potentials p n and p p of neurons and protons from their common 
value po and the pairing gap Ao in the SD channel at 5p = 0. With increasing isospin symmetry, 
i.e., as 5p increases from zero to values of order Ao, a sequence of unconventional phases may 
emerge. One of these is a neutron-proton condensate whose Cooper pairs have non-zero center- 
of-mass (CM) momentum [25} 126] . This phase is the analogue of the Larkin-Ovchinnikov- 
Fulde-Ferrell (LOFF) phase in electronic superconductors [271 128] , Another possibility is phase 
separation into superconducting and normal components, proposed in the context of cold atomic 
gases |29j. At large isospin asymmetry, where SD pairing is strongly suppressed, a BCS-BEC 
crossover may also occur in the isotriplet 1 Sq pairing channel. The ideas of unconventional 
SD pairing and the BCS-BEC crossover in a model of isospin-asymmetric nuclear matter 
where combined recently in Ref. |21j . A phase diagram for superfluid nuclear matter over 
wide ranges of density, temperature, and isospin asymmetry was constructed. A self-consistent 
set of equations, which includes the gap equation and the expressions for the densities of 
constituents (neutrons and protons) was solved allowing for the ordinary BCS state, its low- 




-2.5 -2 -1.5 -1 -0.5 



log 10 (p/p ) 

Figure 1. Phase diagram of dilute nuclear matter in the temperature-density plane for several 
isospin asymmetries a (from Ref. (211). Included are four phases: unpaired phase, BCS (BEC) 
phase, LOFF phase, and PS-BCS (PS-BEC) phase. For each asymmetry there are two tri-critical 
points, one of which is always a Lifshitz point |3Q| . For special values of asymmetry these two 
points degenerate into single tetra-critical point at log(p/po) = —0.22 and T = 2.85 MeV (shown 
by a square dot) for 04 = 0.255. The LOFF phase disappears at the point log(p/po) = —0.65 
and T = (shown by a triangle) for a = 0.62. 



density asymptotic counterpart BEC state and two phases that owe their existence to the isospin 
asymmetry: the phase with a moving condensate (LOFF phase) and the phase where the normal 
fluid and superfluid break down into separate domains. 

The phase diagram found in Ref. |21j is shown in Fig[TJ The results for the BCS phase and 
BCS-BEC crossover are consistent with the earlier studies: one observes a smooth crossover 
to an asymptotic state corresponding to a mixture of a deuteron Bose condensate and a gas 
of excess neutrons. The transition from BCS to BEC is established according the following 
criteria: (i) The average chemical potential p changes its sign from positive to negative values, 
(ii) the coherence length of a Cooper pair becomes comparable to the interparticle distance, i.e., 
£ ~ d ~ p -1 / 3 as conditions change from £ 3> d to £ <C d. 

The nuclear LOFF phase arises as a result of the energetic advantage of translational 
symmetry breaking by the condensate, in which pairs acquire a non-zero CM momentum Q. 
At constant asymmetry, a temperature increase shifts the gap maximum and the free-energy 
minimum of the LOFF phase toward small Q, and at sufficiently high temperature and small 
asymmetry the BCS state is favored over the LOFF phase. This behavior is well understood in 
terms of the phase-space overlap of the Fermi surfaces of neutrons and protons, which (at finite 
asymmetry) increases with temperature and the momentum Q of the Cooper pairs. 

Thus, as the temperature increases, we expect a restoration of the BCS phase and of the 
translational symmetry in the superfluid. Obviously, the same restoration occurs when the 
isospin asymmetry is small enough. The superfluid phase with phase separation (PS) has the 



symmetrical BCS phase as one of its components. The temperature dependence of this phase is 
well established within BCS theory. The second component, which accommodates the neutron 
excess, is a normal Fermi liquid whose low-temperature thermodynamics is controlled by the 
excitations in the narrow strip of width T/ep^ n i p around the Fermi surfaces of neutrons and 
protons. 

The transition to the BEC regime of strongly-coupled neutron-proton pairs, which are 
asymptotically identical with deuterons, occurs at low densities. As already well established, in 
the case of neutron-proton pairing the criteria for the BCS-BEC transition are fulfilled, i.e., fl 
changes sign and the mean distance between the pairs becomes larger than the coherence length 
of the superfluid. 

We now turn to the question of how the BCS-BEC crossover is affected by the existence of 
nuclear LOFF and PS phases at non-zero isospin asymmetries, and conversely how these phases 
evolve in the strongly-coupled regime if the density of the system is decreased. Four different 
phases of matter are present in the phase diagram (Fig. [I]), (i) The unpaired phase is always 
the ground state of matter at sufficiently high temperatures T > T c o, where T c o(p) is the critical 
temperature of the superfluid phase transition at a = 0. (ii) The LOFF phase is the ground 
state in a narrow temperature-density strip at low temperatures and high densities, (iii) The PS 
phase appears at low temperatures and low densities, while the isospin-asymmetric BCS phase 
is the ground state for all densities at intermediate temperatures. In the extreme low-density 
and strong-coupling regime the BCS superfluid phases have two counterparts: the BCS phase 
evolves into the BEC phase of deuterons, whereas the PS-BCS phase evolves into the PS-BEC 
phase, in which the superfluid fraction of matter is a BEC of deuterons. The superfluid-unpaired 
phase transitions and the phase transitions between the superfluid phases are of second order 
(thin solid lines in Fig. [I]), with the exception of the PS-BCS to LOFF transition, which is of 
first order (thick solid lines in Fig. HJ. The BCS-BEC transition and the PS-BCS to PS-BEC 
transition are smooth crossovers. At non-zero isospin asymmetry the phase diagram features two 
tri-critical points where the simpler pairwise phase coexistence terminates and three different 
phases coexist. 

The extreme low-density region of the phase diagram features two crossovers. At intermediate 
temperatures one recovers the well-known BCS-BEC crossover, where the neutron-proton BCS 
condensate transforms smoothly into a BEC gas of deuterons with some excess of neutrons. The 
new ingredient of our phase diagram is the second crossover at low temperatures, where the 
heterogeneous superfluid phase is replaced by a heterogeneous mixture of a phase containing a 
deuteron condensate and a phase containing neutron-rich unpaired nuclear matter. 

What would be the effect of adding somewhat heavier clusters, notably 3 He, 3 H and 4 He, to 
the phases discussed above? At low densities statistical equilibrium suggests that the heavier 
clusters may "eat up" some of the phase-space, therefore their main effect would be to reduce 
the phase space available to pair-correlated particles, which would eventually lead to some shifts 
in the phase diagram without qualitative modifications of its structure or topology. The physics 
might not be as trivial at high densities, for we know that the asymptotically dense phase of 
matter does not contain clusters. The transition to the homogenous phase can be attributed 
to the Pauli-blo eking of the phase space that can be used to form three and four body bound 
states. In other words, the rise of the Fermi surface suppresses the formation of three and 
four-body bound states. It remains an open problem to tackle the interplay of the clusters and 
pair-correlations on the ultimate composition of matter at subnuclear densities. As stated in the 
introduction, the answer is not merely of academic interest. The neutrinos (and more generally 
leptons) transport energy in the supernova processes while moving through this environment; 
its composition, many-body effects, etc remain the key unknowns in the accurate description of 
the neutrino transport. 



3. Neutrino pair-breaking processes 

Next we trasport ourselves in time several weeks past the supernova explosion and the formation 
of a neutron star. During this period of time the star has cooled down to the temperature of 
superfluid phase transition of baryons, which is roughly in the range T c ~ 0.5 MeV. At this 
stage the crust of the star cools predominantly via the electron (e~ ) bremsstrahlung process: 
e~ + A — > e" + A + u + where A stands for a nucleus and v and v for the neutrino and 
antineutrino [3T] . The surface of the star cools by emitting thermal soft X-rays (photons) . Both 
processes are not important sinks of energy up until the star's age t > 10 — 10 5 yr [2J O H]. 
The exact value of the transition to the crust plus surface cooling depends on a number of 
factors, which we will not discuss. While it was widely accepted from the beginning that the 
superfluidity will suppress the neutrino radiation processes on baryons in the core, once it 
becomes superfluid, the less trivial pair-breaking processes where initially neglected, although 
the rates of these processes, computed at one-loop, where available since 1976 [32l [331 [M] . The 
Standard Model weak neutral interactions proceed via vector and axial vector interactions. 
Thus, the neutrino emissivity (phase space integrated rate of neutrino emission) is mainly 
determined by the response functions of matter to weak vector and axial- vector currents. These 
can be computed within the finite temperature Green's functions technique in general. Specific 
applications within the real-time Schwinger-Keldysh formalism can be found in Refs. [331 135j. 

Before discussing these response functions, we pause to recall the many-body developments 
of the formal theory. The polarization tensor in superfluid system involves a re-summation of 
infinite series of particle- hole diagrams. The methods for doing so were developed in the case of 
symmetrical nuclear matter, with the purpose of applications to finite nuclei in Ref. [36]. This 
work is one of the first applications of the Landau Fermi liquid theory (which was designed to 
describe the properties of non-superfluid liquid He 3 ) to superfluid systems. The diagrammatic 
methods of re-summation of particle- hole ladders were developed even earlier [37} [38] . While the 
theory of superfluid Fermi liquids was generalized further by Leggett [39] to finite temperatures 
(in the context of condensed matter theories) , Leggett did so by computing directly the response 
functions and not the vertex functions. In nuclear physics, the theory took a turn that focused 
the entire subsequent work on finite systems (i.e. nuclei), which were treated in the equation of 
motion formalism for second-quantized operators. 

The importance of the vertex corrections for the case of the vector current interactions was 
first pointed out in Ref. [ID], where the authors implemented a polarization tensor adopted from 
condensed matter work (derived not quite in the regime needed for neutron stars). A number 
of subsequent works derived directly the finite temperature vertex functions and polarization 
tensors for neutron and proton components of the core of the star |4"2"1 H3"l 144] . The current 
consensus is that the vector current emissivity is suppressed compared to the one-loop results 
by a factor (vf/c) 4 , where Vf/c ~ 0.1 is the Fermi velocity of the baryons in units of the speed 
of light. At the same time it was established that the axial vector interactions are adequately 
represented by the bare vertices and one-loop polarization tensors. Explicitly, the axial vector 
emissivity given by |34[ |4"2] 

e„= 2 -^^u(0)vf 2 T 7 I„ I v = f jT 'dy-^=f F {zy)\ (1) 

where Gf = 1.166 x 10~ 5 GeV~ 2 is the Fermi weak coupling constant, gA = 1-25 is 
(unquenched) axial vector coupling constant, v(0) is the density of states at the Fermi surface, 
T is the temperature, z = A(T)/T is the ratio of the pairing gap to the temperature, and 
Jf( x ) = [1 + exp(x)]" 1 is the Fermi distribution function. The emissivity of the axial vector 
current is of the order (vf/c) 2 and, therefore, is the dominant channel of the energy loss. 



4. Axion emission from compact stars 

The strong sector of the Standard Model may feature a CP violating interaction, which arises 
due to a topological interaction term in the QCD Lagrangian |45j 



•a 



(2) 



32vr 2 » v 



where F* v = d^A u - d v A^ + gf abc A^ b A uc is the gluon field strength tensor, F£, = e^ Xp F xpa /2, 
j?abc are ^Yi e structure constants of SU(3) group, 9 is the parameter which parametrizes the 
non-perturbative vacuum states of QCD \9) = J2 n ex P( — in9)\n), where n is the winding number 
characterizing each distinct state of QCD, which is not connected to another by any gauge 
transformation. The QCD action changes by 2ir under the shift 9^-9 + 2ir, i.e., 9 is a 
periodic function with a period of 2ir. In presence of quarks the physical parameter is not 
9, but 9 = 9 + argdetm 9 , where m q is the matrix of quark masses. Experimentally, the upper 
bound on the value of this parameter is 9 < 10~ 10 , which is based on the measurements of the 
electric dipole moment of neutron d n < 6.3 • 10 -26 e cm |46| . The smallness of 9 is the strong 
CP problem: the Standard Model does not provide any explanation on why this number should 
not be of order unity. 

An elegant solution to the strong CP problem is provided by the Peccei-Quinn mechanism [47, 
|4"H1 149] . This solution amounts to introducing a global U(1)pq symmetry, which adds an 
additional anomaly term to the QCD action proportional to the axion field a. This term acts as 
a potential for the axion field and gives rise to an expectation value of the axion field (a) ~ —9. 
The physical axion field is then a — (a), so that the undesirable 9 term in the action is replaced by 
the physical axion field. The axion is the Nambu-Goldstone boson of the Peccei-Quinn U{\)pq 
symmetry breaking |48[ l4~9] , and its effective Lagrangian has the form 



where the second term describes the coupling of the axion to fermion fields (-0) of the Standard 



There are ongoing experimental searches for the axion. Cosmology and astrophysics provide 
strong complementary constraints. Because axions can be effectively produced in the interiors of 
stars they act as an additional sink of energy. The requirements that the energy loss from a star 
is consistent with the astrophysical observations place lower bounds on the coupling of axions to 
the standard model particles, and hence on the Peccei-Quinn symmetry breaking scale I5U| . 
The latter limit translates into an upper limit on the axion mass. Such arguments have been 
applied to the physics of supernova explosions [511 [52l [531 EH [55] and white dwarfs [56] . In 
the case of supernova explosions the dominant energy loss process is the emission of an axion 
in the nucleon (n) bremsstrahlung n + n — > n + n + a. The same process was considered 
earlier by Iwamoto as a cooling mechanism for mature neutron stars, i.e., neutron stars with 
core temperature in the range 10 8 — 10 9 K |57j . The implications of the axion emission by 
the modified Urea and nucleon bremsstrahlung, as calculated in Ref. [57], were briefly studied 
in Ref. [58]. However, as discussed in the previous section the pair-breaking processes are the 
dominant energy loss channels for temperatures below the critical temperature. Therefore, to 
obtain bounds on the properties of axions from the cooling of compact stars, the rate of the axion 
cooling is required. This was computed recently in Ref. [6] in the case of 5-wave superfluid. 

We now briefly outline this calculation. The energy radiated per unit time in axions (axion 
emissivity) is given by the phase-space integral over the probability of the emission process 




(3) 



Model. 




(4) 



x-~ m ~-^C x x*- ^ jh~~~~~x 



Figure 2. The two diagrams contributing to the polarization tensor of baryonic matter, 
which defines the axion emissivity. The "normal" baryon propagators for particles (holes) are 
shown by single-arrowed lines directed from left to right (right to left). The double arrowed 
lines correspond to the "anomalous" propagators with two incoming or outgoing arrows. The 
horizontal dashed lines represent the axion a. 



where q and u are the axion momentum and energy, Ii a iu {q) is the polarization tensor of the 
superfluid baryonic matter, <?b(w) is the Bose distribution function, f a is the axion-baryon 
coupling strength. The polarization tensor of a superfluid obtains contributions from four 
distinct diagrams that can be formed from the normal and anomalous propagators with four 
distinct effective vertices [JTJ S3 H3 HI] . However, for the axial vector perturbations the vertices 
are not renormalized in the medium and, therefore, one proceeds with the bare vertices, in which 
case the number of the distinct contributions to the polarization tensor reduces to a sum of two 
admissible bare loops (see Fig. [2]). The axion emissivity obtained from Eq. (Jl|) is given by 

e a = —f- 2 v(Q)v}T 5 I a , I a = z 5 J^dy-j^=f F {zy) 2 . (5) 

The T 5 scaling of the emissivity is understood as follows. The integration over the phase space 
of neutrons carries a power of T, since for degenerate neutrons the phase-space integrals are 
confined to a narrow strip around the Fermi surface of thickness T. The axion is emitted 
thermally and being relativistic contributes a factor T 3 to the emissivity. The one power of T 
from the energy of the axion and the inverse one power of T from the energy conserving delta 
function cancel. The transition matrix element is proportional to the combinations of u p and v p 
amplitudes, which are dimensionless, but contain implicit temperature dependence due to the 
temperature dependence of the gap function. This dependence is not manifest in Eq. ©, i.e., 
was absorbed in the definition of the integral I a . Thus, the explicit temperature dependence 
of the axion emission rate Eq. ([5]) is T 5 . As discussed in Sec. [3] at temperatures of order the 
critical temperature T c ~ 10 9 K the superfluid cools primarily by emission of neutrinos via the 
pair-breaking processes driven by the axial- vector currents (we continue to assume that potential 
fast cooling via direct Urea processes is prohibited). In a first approximation one may require 
that the axion luminosity does not exceed the neutrino luminosity, i.e., 

± = 10 " 2 k~ 59 - 2 r(z)<1 (6) 

where r(z) = z 2 (I a /I u ). Not far from the critical temperature A(T) ~ 3.06T c a/1 — T/T c , which 
translates into z = 3.06 1^ 1 \/l — t, where t = T/T c . Numerical evaluations of the integrals 
provides the following values r(0.5) = 0.07, r(l) = 0.26, r(2) = 0.6 and asymptotically r(z) — > 1 
for z> 1, Noting that r(z) < 1 , one finally obtains 



f a > 5.92 x 10 9 GeV 



0.1 MeV 



A(T) 

which translates into an upper bound on the axion mass 

, /l0 10 GeV\ , f ACT) 

m a = 0.62 x lO- 3 eV < 1.05 x 10~ 3 eV - - \ \ 

V f a I L0.1 MeV 



(7) 



(8) 



The bound Eq. (|7|) can be written in terms of the critical temperature by noting that A(T) ~ T c 
in the temperature range 0.5 < t < 1 of interest. 

The neutrino cooling era of compact stars, which spans the time-period t < 10 4 — 10 5 yr after 
their birth in supernova explosions is a sensitive probe of the particle physics of their interiors. 
If one assumes that there are no rapid channels of cooling in neutron stars, i.e., deconfined 
quarks, above Urea threshold fractions of protons or hyperons (all of which lead to a rapid 
Urea cooling), then neutron stars cool primarily by neutrino emission in Cooper pair-breaking 
processes in baryonic superfluids. If, however, axions exist in Nature, the neutron stars must cool 
via axion emission in Cooper pair-breaking processes, whose axion emission rate scales as T 5 . 
This scaling differs from the T 7 scaling of the counterpart neutrino processes. The difference 
arises from the different phase spaces required for the pairs of neutrinos and the axion and 
is independent of the baryonic polarization tensor. Note also that the rate of axion emission 
from a P-wave superfluid will differ from the S-wave rate, derived above, by a factor 0(1) and, 
therefore, will not change quantitatively the obtained bounds on the axion parameters. 

Similar bounds to those quoted above were obtained previously by Iwamoto [57] 
(//10 10 GeV > 0.3) from a comparison of the rates of axion bremsstrahlung and modified Urea 
neutrino emission by mature neutron stars, and by Umeda et al [58J (//10 10 GeV > 0.1 — 0.2) 
from fits of cooling simulations to the PSR 0656+14 dats0- The lower bound on f a derived in 
Ref. [6] is somewhat larger than the one that follows form the requirement that the axions do not 
"drain" too much energy from supernova process so that it fails [5[ [5TJ [52j [531 [Ml [55] . Note the 
dependence of the bound Eq. ([7]) on the pairing gap. This is an ingredient, which originates from 
superfluidity of cold neutron stars, that does not appear in other bounds on axion parameters. 
Because, the magnitude and density dependence of the gap are not well-known (for a review 
see Ref. |59j) there remains enough room for speculations on the impact of the axions on the 
cooling of compact stars. 

5. Perspectives 

Although the equation of state and composition of the dilute and warm nuclear matter have been 
studied for several decades, there remain a number of unsettled issues related to the complex 
many-body character of this system. The interplay of the pair-correlations and clustering is 
one of the challenging problems. As we have seen the phase structure of the matter at finite 
isospin becomes more complicated due to the emergence of novel superconducting phases. It is 
very likely that along with the deuteron condensate and its BCS counterpart a BEC of alpha 
particles will coexist with these phases. Thus, one needs to work out the physics of a mixtures 
of several superfluids phases in this context. How these new features will affect the neutrino 
transport in supernovas remains an open question. 

Neutron star cooling offers a new playground for the studies of beyond Standard Model 
physics. The example presented above shows that with microscopic rates of the axion emission 
from superfluid phases at hand one can place useful bounds on the axion properties. Axions 
offer a unique channel of rapid energy loss by medium to low mass compact stars, where the 
central densities are below the direct Urea thresholds and/or below the density of deconfinement 
into quark matter. Thus, the long-standing paradigm that the light and medium mass neutron 
stars cannot cool rapidly will not hold should the axions exist in Nature and couple to matter 
strong enough to cool a compact star faster than the Standard Model neutrinos. 

Note that Eq. (9) of this work is incorrect, which could be the reason why the limits on the axion's mass 
reported in their Figs. 2-4 are by an order of magnitude larger compared to those quoted in Ref. [6]. 



Acknowledgements 

I gratefully acknowledge the discussions and collaboration with J. W. Clark, X.-G. Huang, J. 
Keller, and M. Stein on the topics presented in this lecture. 



References 



A. Burrows, arXiv: 1210.4921 [astro-ph.SR]. 

D. G. Yakovlev, O. Y. Gnedin, A. D. Kaminker and A. Y. Potekhin, AIP Conf. Proc. 983, 379 (2008). 

D. Page, in Neutron Stars and Pulsars, Astrophys. Space Sci. Library, vol. 357, ed. W. Becker, (Springer- 
Verlag, 2009), p. 247. 

A. Sedrakian, Prog. Part. Nucl. Pbys. 58, 168 (2007). 

G. G. Raffelt, Lect. Notes Phys. 741, 51 (2008). 

J. Keller and A. Sedrakian, Nucl. Phys. A 897, 62 (2013). 

K. Sumiyoshi and G. Ropke, Phys. Rev. C77, 055804 (2008). 

S. Heckel, P. P. Schneider and A. Sedrakian, Phys. Rev. C80, 015805 (2009). 

S. Typel, G. Ropke, T. Klahn, D. Blaschke and H. H. Wolter, Phys. Rev. C81, 015803 (2010). 

M. Hempel and J. Schaffner-Bielich, Nucl. Phys. A837, 210 (2010). 

A. R. Raduta and F. Gulminelli, Phys. Rev. C82, 065801 (2010). 

M. Oertel, A. F. Fantina and J. Novak, Phys. Rev. C85, 055806 (2012). 

F. Gulminelli and A. R. Raduta, Phys. Rev. C85, 025803 (2012). 

T. Aim, B. L. Friman, G. Ropke and H. Schulz, Nucl. Phys. A551, 45 (1993). 

M. Baldo, U. Lombardo and P. Schuck, Phys. Rev. C52, 975 (1995). 

A. Sedrakian and J. W. Clark, Phys. Rev. C73, 035803 (2006). 

S. Mao, X. Huang and P. Zhuang, Phys. Rev. C79, 034304 (2009). 

X. -G. Huang, Phys. Rev. C81, 034007 (2010). 

M. Jin, M. Urban and P. Schuck, Phys. Rev. C82, 024911 (2010). 

G. Ropke, N. -U. Bastian, D. Blaschke, T. Klahn, S. Typel and H. H. Wolter. [arXiv: 1209.02121 [nucl-th], 
M. Stein, X. -G. Huang, A. Sedrakian and J. W. Clark, Phys. Rev. C in press. [axXlvTl 208.0123 [nucl-th]. 
P. Nozieres and S. Schmitt-Rink, J. Low. Temp. Phys. 59, 195 (1985). 

A. Sedrakian and U. Lombardo, Phys. Rev. Lett. 84, 602 (2000). 

U. Lombardo, P. Nozieres, P. Schuck, H. J. Schulze and A. Sedrakian, Phys. Rev. C64, 064314 (2001). 
A. Sedrakian, Phys. Rev. C63, 025801 (2001). 

H. Miither and A. Sedrakian, Phys. Rev. C67, 015802 (2003). 

A. I. Larkin and Y. N. Ovchinnikov. Zh. Eksp. Teor. Fiz., 47 762 (1965). 
P. Fulde and R. A. Ferrell, Phys. Rev. 135, 550 (1965). 

P. F. Bedaque, H. Caldas and G. Rupak, Phys. Rev. Lett. 91, 247002 (2003). 

P. Chaikin and T. Lubensky, Principles of Condensed Matter Physics (Cambridge University Press, 1995). 
S. L. Shapiro and S. A. Tuekolsky, Black Holes, White Dwarfs, and Neutron Stars (John. Wiley & Sons, 
New York, 1983). 

E. Flowers, M. Ruderman, P. Sutherland, Astrophys. J. 205, 541 (1976). 

D. N. Voskresensky, A. V. Senatorov, Sov. J. Nucl. Phys. 45, 657 (1987) [Yad. Fiz. 45, 411 (1987)]. 
A. D. Kaminker, P. Haensel, D. G. Yakovlev, Astron. Astrophys. 345, L14-L16 (1999). 

A. Sedrakian, A. Dieperink, Phys. Lett. B463, 145-152 (1999). Phys. Rev. D62, 083002 (2000). 

A. I. Larkin and A. B. Migdal, Sov. Phys. JETP 17, 1146 (1963); A. B. Migdal, Theory of Finite Fermi 

Systems and Applications to Atomic Nuclei (Interscience, London, 1967). 
A. A. Abrikosov, L. P. Gorkov, Sov. Phys. JETP 8, 1090 (1959); Sov. Phys. JETP 9, 220 (1959). 
A. A. Abrikosov, L. P. Gorkov, and I. E. Dzyaloshinski, Methods of quantum field theory in statistical physics, 

(Dover, New York, 1975). 
A. J. Leggett, Phys. Rev. 147, 119 (1966). 
L. B. Leinson and A. Perez, Phys. Lett. B 638, 114 (2006). 
A. Sedrakian, H. Miither, P. Schuck, Phys. Rev. C76, 055805 (2007). 

E. E. Kolomeitsev, D. N. Voskresensky, Phys. Rev. C77, 065808 (2008); Phys. Rev. C 81, 065801 (2010). 
A. W. Steiner, S. Reddy, Phys. Rev. C79, 015802 (2009). 

A. Sedrakian, Phys. Rev. C86, 025803 (2012). 

G. 't Hooft, Phys. Rev. Lett. 37, 8 (1976). 

C. A. Baker et al, Phys. Rev. Lett. 97, 131801 (2006). 

R. D. Peccei and H. R. Quinn, Phys. Rev. Lett. 38, 1440 (1977). 

S. Weinberg, Phys. Rev. Lett. 40, 223 (1978). 

F. Wilczek, Phys. Rev. Lett. 40, 279 (1978). 



[50] G. G. Raffelt, J. Redondo and N. V. Maira, Phys. Rev. D 84, 103008 (2011). 
[51] R. P. Brinkmann and M. S. Turner, Phys. Rev. D 38, 2338 (1988). 
[52] A. Burrows, M. S. Turner and R. P. Brinkmann, Phys. Rev. D 39, 1020 (1989). 
[53] G. Raffelt and D. Seckel, Phys. Rev. D 52, 1780 (1995). 

[54] H.-T. Janka, W. Keil, G. Raffelt and D. Seckel, Phys. Rev. Lett. 76, 2621 (1996). 
[55] C. Hanhart, D. R. Phillips and S. Reddy, Phys. Lett. B 499, 9 (2001). 

[56] A. H. Corsico, L. G. Althaus, M. M. M. Bertolami, A. D. Romero, E. Garcia-Berro, J. Isern and S. O. Kepler, 

Mon. Not. R. Astron. Soc., 424, 2792 (2012). 
[57] N. Iwamoto, Phys. Rev. Lett. 53, 1198 (1984). 

[58] H. Umeda, N. Iwamoto, S. Tsuruta, L. Qin and K. Nomoto, in Proceedings of the International Conference 
on Neutron Stars and Pulsars, edited by N. Shibazaki et al., Frontiers science series no. 24, (Universal 
Academy Press, 1998), p. 213. [astro-ph/9806337| . 

[59] A. Sedrakian and J. W. Clark, in "Pairing in Fermionic Systems: Basic Concepts and Modern Applications" , 
edited by A. Sedrakia n, J. W. Clark and M. Alford, (World Scientific, Singapore, 2006), p. 135. 
arXiv:nucl-th/0607028j .